<<

Masatsugu Sei Suzuki Department of Physics, SUNY at Binghamton (Date: March 04, 2017)

In classical mechanics, identical particles (such as electrons) do not lose their "individuality", despite the identity of their physical properties. In , the situation is entirely different. Because of the Heisenberg's principle of uncertainty, the concept of the path of an electron ceases to have any meaning. Thus, there is in principle no possibility of separately following each of a number of similar particles and thereby distinguishing them. In quantum mechanics, identical particles entirely lose their individuality. The principle of the in-distinguishability of similar particles plays a fundamental part in the quantum theory of system composed of identical particles. Here we start to consider a system of only two particles. Because of the identity of the particles, the states of the system obtained from each other by interchanging the two particles must be completely equivalent physically.

1 System of particles 1 and 2

k' , k"

  k' k"   k" k' a 1 2 b 1 2

Even though the two particles are indistinguishable, mathematically  a and  b are distinct kets for

k'  k" . In fact we have  a  b  0 . Suppose we make a measurement on the two particle system.

k' : state of one particle and k" : state of the other particle.

We do not know a priori whether the state ket is   k' k" or   k" k' or -for that matter- any a 1 2 b 1 2 linear combination of the two: ca  a  cb  b .

2 Exchange degeneracy A specification of the eigenvalue of a complete set of observables does not completely determine the state ket.

Mathematics of permutation symmetry:

Pˆ   Pˆ k' k"  k" k'   . 12 a 12 1 2 1 2 b

Clearly

ˆ ˆ ˆ 2 P21  P12 , and P12 1.

ˆ Under P12 , particle 1 having k' becomes particle 1 having k" ; particle 2 having k" becomes particle 1 having k' . In other words, it has the effect of interchanging 1 and 2.

((Note))

Identical particles 1 9/3/2017 Matrix element of Pˆ  Pˆ in terms of   k' k" and   k" k' 21 12 a 1 2 b 1 2

Pˆ   Pˆ k' k"  k" k'   , 12 a 12 1 2 1 2 b

Pˆ   Pˆ k" k'  k' k"   . 12 b 12 1 2 1 2 a

ˆ ˆ Matrix element of P21  P12

ˆ 0 1 P12    1 0 or

 a  b

 a 0 1

 b 1 0

ˆ Eigensystem[ P12 ]

 = 1 (symmetric):

P12  symm   symm

1 1   ( k' k"  k" k' )  (1ˆ  Pˆ ) k' k" symm 2 1 2 1 2 2 12 1 2 where the symmetrizer is defined by

1 Sˆ  (1ˆ  Pˆ ) 2 12

= -1 (anti-symmetric)

P12  anti   anti

1 1   ( k' k"  k" k' )  (1ˆ  Pˆ ) k' k" anti 2 1 2 1 2 2 12 1 2 where the antisymmetrizer is defined by

1 Aˆ  (1ˆ  Pˆ ) 2 12

Identical particles 2 9/3/2017 Our consideration can be extended to a system made up of many identical particles. A transposition is a permutation which simply exchange the role of two of the particles, without touching others.

Pˆ k' k" ..... k i k i1 ..... k j .....  k' k" ..... k j k i1 ..... k i ..... ij 1 2 i i1 j 1 2 i i1 j

ˆ ˆ  ˆ The transposition operators Pij are Hermitian ( Pij  Pij )

ˆ 2 ˆ Pij  1

ˆ ˆ So that Pij is an unitary operator. The allowed eigenvalues of Pij are ±1. It is important to note, however, that in general

ˆ ˆ [Pij , Pkl ]  0.

(i) The first example

ˆ Now we consider a permutation operator P123 for

Pˆ k' k" k"'  k' k" k"'  k"' k' k" 123 1 2 3 2 3 1 1 2 3 for the system of 3 identical particles ( k' k" k''' ) 1 2 3

123 P    123 231

This means replacement of 12, 2 3, 3 1.

123 123213 P       P P 123 231 213231 12 13

Quantum mechanically this is not correct. The correct one is

ˆ ˆ ˆ P123  P13 P12

(ii) The second example

123 12 3132 P       P P 123 231 132231 23 12 or quantum mechanically

ˆ ˆ ˆ P123  P12P23

(iii) The third example

Identical particles 3 9/3/2017

123 123321 P       P P 132 312 321312 13 12 or quantum mechanically

ˆ ˆ ˆ P132  P12 P13

((Proof))

Pˆ Pˆ k' k" k'''  Pˆ k''' k" k'  k" k''' k' 12 13 1 2 3 12 1 2 3 1 2 3 and

Pˆ k' k" k'''  k' k" k'''  k" k''' k' 132 1 2 3 3 1 2 1 2 3

ˆ ˆ ˆ Therefore we have P132  P12 P13 .

Any permutation operators can be broken into a product of transposition operators.

ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ 2 P132  P23 P12  P13 P23  P12 P13  P23 P12 P23  ...

The decomposition is not unique. However, for a given permutation, it can be shown that the parity of the number of transposition into which it can be broken down is always the same: it is called the parity of the permutation.

ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ 2 P132  P23P12  P13P23  P12P13  P23P12P23 even permutation

ˆ ˆ ˆ ˆ ˆ P123  P12 P23  P13 P12 : even permutation

ˆ P23 odd permutation

ˆ P12 odd permutation

ˆ P31 odd permutation

((Note))

ˆ ˆ ˆ P132  P321  P213

3 Symmetrizer and antisymmetrizer Now we consider the two Hermitian operators

ˆ 1 ˆ S   P : symmetrizer N! 

Identical particles 4 9/3/2017 ˆ 1 ˆ A   P : antisymmetrizer N! 

ˆ ˆ where  =1 if P is an even permutation and  = -1 if P is an odd permutation.

Sˆ   Sˆ and Aˆ   Aˆ .

((Theorem-1)) ˆ If P 0 is an arbitrary permutation operator, we have

ˆ ˆ ˆ ˆ ˆ P 0 S  SP 0  S

ˆ ˆ ˆ ˆ P 0 A  AP  0   0 A (1)

((Proof)) This is due to the fact that

ˆ ˆ ˆ P 0P  P such that

  0  or

2 0    0   

Consequently

ˆ ˆ 1 ˆ ˆ 1 ˆ ˆ P 0S   P 0P   P  S N!  N! 

ˆ ˆ 1 ˆ ˆ 1 ˆ ˆ P 0 A   P 0 P   0   P   0 A N!  N! 

From Eq.(1), we see the following theorem

((Theorem-2))

Sˆ 2  Sˆ

Aˆ 2  Aˆ and

AˆSˆ  SˆAˆ  0 Identical particles 5 9/3/2017

((Proof))

ˆ 2 1 ˆ ˆ 1 ˆ ˆ S   P S   S  S N!  N! 

ˆ 2 1 ˆ ˆ 1 2 ˆ ˆ A   P A   A  A N!  N! 

ˆ ˆ ˆ 1 ˆ ˆ 1 ˆ AS  A   P S  S  0 N!  N!  since half the  are equal to 1 and half the  equal to -1. We also note that

1 2 1   1  1 N!  N! 

Sˆ and Aˆ are therefore projectors. Their action on any ket  of the state space yields a completely symmetric or completely antisymmetric ket.

ˆ ˆ ˆ P 0S   S 

Pˆ Aˆ    Aˆ   0  0

ˆ  S  S 

ˆ  A  A

((Example))

For N = 3,

1 Sˆ  [1ˆ  Pˆ  Pˆ  Pˆ  Pˆ  Pˆ ] 6 12 23 31 123 132 and

1 Aˆ  [1ˆ  Pˆ  Pˆ  Pˆ  Pˆ  Pˆ ] 6 12 23 31 123 132 where

ˆ ˆ ˆ ˆ ˆ ˆ P123  P12P23 , P132  P12P13

Identical particles 6 9/3/2017 1 Sˆ  Aˆ  (1ˆ  Pˆ  Pˆ )  1ˆ 3 123 132

4 Symmetrization postulate The system containing N identical particles are either totally symmetrical under the interchange of any pair (boson), or totally antisymmetrical under the interchange of any pair ().

ˆ Pij  N ,B   N ,B ,

ˆ Pij  N ,F   N ,F ,

where  N ,B is the eigenket of N identical boson systems and  N ,F is the eigenket of N identical fermion systems.

((Note)) It is an empirical fact that a mixed symmetry does not occur.

Even more remarkable is that there is a connection between the spin of a particle and the statistics obeyed by it:

Half-integer spin particles are fermion, while integer-spin particles are bosons.

5 Pauli exclusion principle Wolfgang Ernst Pauli (April 25, 1900 – December 15, 1958) was an Austrian theoretical physicist and one of the pioneers of quantum physics. In 1945, after being nominated by Albert Einstein, he received the Nobel Prize in Physics for his "decisive contribution through his discovery of a new law of Nature, the exclusion principle or Pauli principle," involving spin theory, underpinning the structure of matter and the whole of chemistry.

http://en.wikipedia.org/wiki/Wolfgang_Pauli

Electron is a fermion. No two electrons can occupy the same state. We discuss the framatic difference between and bosons. Let us consider two particles. Each of which can occupy only two states k' and k" . For a system of two fermions, we have no choice

Identical particles 7 9/3/2017

1 1 k' k' [ k' k"  k' k" ]  1 2 1 2 2 1 k'' k" 2 2 1 2

For bosons, there are three states.

k' k' , 1 2

k" k" , 1 2

1 ( k' k"  k" k' ) . 2 1 2 1 2

In contrast, for “classical particles” satisfying Maxwell-Boltzmann (M-B) statitics with no restriction on symmetry, we have altogether four independent states.

k' k" , k" k' , k' k' and k" k" 1 2 1 2 1 2 1 2

We see that in the fermion case, it is impossible for both particles to occupy the same state.

6 Transformation of observables by permutation For simplicity, we consider a specific case where the two particle state ket is completely specified by the eigenvalues of a single observable Aˆ for each of the particle.

Aˆ a' a"  a' a' a" 1 1 2 1 2 and

Aˆ a' a"  a" a' a" 2 1 2 1 2

Since

Pˆ Aˆ Pˆ 1Pˆ a' a"  Pˆ Aˆ Pˆ 1 a" a' 12 1 12 12 1 2 12 1 12 1 2 or

Pˆ Aˆ Pˆ 1Pˆ a' a"  Pˆ Aˆ a' a" 12 1 12 12 1 2 12 1 1 2  a' Pˆ a' a" 12 1 2  a' a" a' 1 2  Aˆ a" a' 2 1 2  Aˆ Pˆ a' a" 2 12 1 2 we obtain Identical particles 8 9/3/2017

ˆ ˆ ˆ 1 ˆ P12 A1 P12  A2 .

Similarly, we have

ˆ ˆ ˆ 1 ˆ P21 A2 P21  A1 .

ˆ It follows that P12 must change the particle label of observables.

ˆ ˆ ˆ ˆ There are also observables, such as A1  B2 , A1 B2 , which involve both indices simultaneously.

ˆ ˆ ˆ ˆ 1 ˆ ˆ P12 (A1  B2 )P12  A2  B1

ˆ ˆ ˆ ˆ 1 ˆ ˆ ˆ 1 ˆ ˆ ˆ 1 ˆ ˆ P12 A1 B2 P12  P12 A1 P12 P12 B2 P12  A2 B1

These results can be generalized to all observables which can be expressed in terms of observables which can ˆ ˆ ˆ be expressed in terms of observables of the type of A1 and B2 , to be denoted by O(1,2) .

ˆ ˆ ˆ 1 ˆ P12 O(1, 2 )P12  O(2,1) where Oˆ (2,1) is the observable obtained from Oˆ (1,2) by exchanging indices 1 and 2 throughout.

ˆ Os (1,2) is said to be symmetric if

ˆ ˆ Os (1, 2)  Os (2,1) or

ˆ ˆ [Os (1,2),P12 ]  0

Symbolic observables commute with the permutation operator.

ˆ In general. the observables Os (1,2,3,..., N) which are completely symmetric under exchange of indices 1, 2, ..., N commute with all the transposition operators, and with all the permutation operators

7 Example Let us now consider a Hamiltonian of a system of two identical particles.

1 1 Hˆ  pˆ 2  pˆ 2 V ( rˆ  rˆ ) V (rˆ ) V (rˆ ) 2m 1 2m 2 pair 1 2 ext 1 ext 2

Clearly we have

ˆ ˆ ˆ 1 ˆ P12 H P12  H or

Identical particles 9 9/3/2017

ˆ ˆ [P12, ,H ]  0

ˆ ˆ 2 ˆ P12 is a constant of the motion. Since P12 1, the eigenvalue of P12 allowed are ±1.

Hˆ   E 

ˆ P12    

ˆ 2 ˆ 2 P12   P12       or

= ±1.

It therefore follows that if the two-particle state ket is symmteric (antisymmettric) to start with, it remains so at all times.

(i) N = 2 case We can define the symmetrizer and antisymmetrtizer as follows.

1 1 Sˆ  (1 Pˆ ) Aˆ  (1 Pˆ ) 2 12 2 12

Sˆ  Aˆ  1ˆ

1 1 1 1 Sˆ 2  (1ˆ  Pˆ ) (1ˆ  Pˆ )  (1ˆ  2Pˆ 1ˆ)  (1ˆ  Pˆ )  Sˆ 2 12 2 12 4 12 2 12

1 1 1 1 Aˆ 2  (1ˆ  Pˆ ) (1ˆ  Pˆ )  (1ˆ  2Pˆ 1ˆ)  (1ˆ  Pˆ )  Aˆ 2 12 2 12 4 12 2 12

Then we have

1  S  ( k' k"  k" k' ) 2 1 2 1 2 and

1  A  ( k' k"  k" k' ) 2 1 2 1 2 where

1 1 Sˆ k' k"  (1 Pˆ ) k' k"  ( k' k"  k" k' ) 1 2 2 12 1 2 2 1 2 1 2 Identical particles 10 9/3/2017

1 1 Aˆ k' k"  (1 Pˆ ) k' k"  ( k' k"  k" k' ) 1 2 2 12 1 2 2 1 2 1 2

(ii) N = 3 Cases

1 Sˆ  (1  Pˆ  Pˆ  Pˆ  Pˆ  Pˆ ) 6 12 23 31 123 132

1 Aˆ  (1  Pˆ  Pˆ  Pˆ  Pˆ  Pˆ ) 6 12 23 31 123 132

1 Sˆ  Aˆ  (1 Pˆ  Pˆ )  1 3 123 132

k' k" k''' 1 1 1 1 Aˆ k' k" k'''  k' k" k''' 1 2 3 3! 2 2 2 k' k" k''' 3 3 3

Slater Aˆ k' k" k''' is zero if two of individual states coincide. We obtain Pauli’s exclusion principle. 1 2 3

8 Generalized method (Tomonaga) We now consider a system consisting of many spins.

ˆ ˆ ˆ ˆ ˆ S  S1  S2  S3  ...  SN

ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ ˆ S  S  (S1  S2  S3  ...  SN )  (S1  S2  S3  ...  SN ) or

N 2 N 1 ˆ 2 ˆ 2 ˆ ˆ  ˆ 2 2 ˆ ˆ S   Sn  2(Sn  Sn' )  σn   (σn  σn' ) n1 nn' 4 n1 2 nn' or

Sˆ 2 3N 1 ˆ ˆ ˆ 2  1 (σn  σn' )  4 2 nn'

Here we define an operator

ˆ 2 ˆ O   Pnn' N(N 1) nn'

(i) Oˆ is Hermitian Identical particles 11 9/3/2017

((Proof))

ˆ  2 ˆ  2 ˆ ˆ O   Pnn'   Pnn'  O N(N 1) nn' N(N 1) nn'

(ii)

[Pˆ,Oˆ]  0.

As shown in the Appendix,

PˆPˆ  Pˆ Pˆ nn' rr rn'

where the pair (n, n') corresponds to the pair (rn , rn' ) in one-to one.

ˆ ˆ ˆ ˆ P Pnn'   Pnn'P nn. nn.

______We assume that

1 Pˆ  (1ˆ  σˆ  σˆ ) (Dirac ) nn' 2 n n'

2 1 2 1 N(N 1) 1 ˆ ˆ ˆ ˆ ˆ ˆ ˆ O   (1 σn  σn' )  [ 1 σn  σn' ] N(N 1) nn' 2 N(N 1) 2 2 2 nn' or

1 2 ˆ ˆ ˆ ˆ O  [1 σn  σn' ] 2 N(N 1) nn'

Using the relation,

Sˆ 2 3N 1 ˆ ˆ ˆ 2  1 (σn  σn' ) ,  4 2 nn' we get

ˆ 1 ˆ 2 2 ˆ 2 3N 1 N  4 ˆ 4 1 ˆ 2 O  [1 ( 2 S  )]  [ 1 2 S ] 2 N(N 1)  2 2 (N 1) N(N 1) 

ˆ 2 ˆ ˆ 2 2 ˆ [S ,O]  0 . When the eigenvalue of S is given by  S(S 1 ) , the eigenvalue of O is equal to

Identical particles 12 9/3/2017 1 N  4 4S(S 1)   [  ] 2 N 1 N(N 1)

The eigenvalue of Oˆ , () , specifies the symmetry.

Oˆ S  S leading to the eigenvalue  1 for the symmetric state

Oˆ A   A leading to the eigenvalue    1 for the antisymmetric state.

(i) For N = 2,

D1/2 x D1/2 = D1 + D0

1   [2  2S(S 1)]  1 S(S 1) 2

When S = 1,  = 1 (symmetric). When S = 0,  = -1 (anti-symmetric).

(ii) For N = 3,

D1/2 x D1/2 x D1/2 = D3/2 + 2 D1/2

1 1 2   [  S(S 1)] 2 2 3

When S = 3/2,  = 1 (symmetric). When S = 1/2,  = 0.

(1) j = 3/2 (symmetric)

3 3 ,   (1) (2) (3) 2 2

3 1 (1) (2) (3)   (1)(2)(3)  (1) (2)(3) ,  2 2 3

3 1  (1)(2)(3) (1) (2)(3) (1)(2) (3) ,  2 2 3

3 3 ,  (1)(2)(3) 2 2

______(ii) j = 1/2.

Identical particles 13 9/3/2017 1 1 (1) (2) (3)   (1)(2) (3)  2 (1) (2)(3) ,  2 2 6

1 1   (1)(2)(3)  2(1)(2) (3) (1) (2)(3) ,  2 2 6

______(iii) j = 1/2

1 1 (1) (2) (3)   (1)(2) (3) [(1) (2))   (1)(2)] (3) ,   2 2 2 2

1 1 (1) (2)(3)   (1)(2)(3) [(1) (2)   (1)(2)](3) ,   2 2 2 2

(iii) For N = 4,

D1/2 x D1/2 x D1/2 x D1/2 = D2 + 3D1 + 2D0

S(S 1)   6

For S = 2,  = 1 (symmetric). For S = 1,  = 1/3. For S = 0,  = 0. (antisymmetric)

9. Summary: fermion and boson The Hamiltonian H of the system (in the absence of a magnetic field) does not contain the spin operators, and hence, when it is applied to the , it has no effect on the spin variables. The wave function of the system of particles can be written in the form of product,

   space spin ,

where  space depends only on the coordinate of the particles and spin only on their spins.    space spin should be anti-symmetric since electrons are Fermions. The system containing N identical particles are either totally symmetrical under the interchange of any pair (boson), or totally antisymmetrical under the interchange of any pair (fermion).

ˆ Pij  N ,B   N ,B ,

ˆ Pij  N ,F   N ,F ,

where  N ,B is the eigenket of N identical boson systems and  N ,F is the eigenket of N identical fermion systms. Identical particles 14 9/3/2017

((Note)) It is an empirical fact that a mixed symmetry does not occur.

Even more remarkable is that there is a connection between the spin of a particle and the statistics obeyed by it:

Half-integer spin particles are fermion, obeying the Fermi-Dirac statistic, while integer-spin particles are bosons, obeying the Bose-Einstein statistics.

10. Ground state for systems with many electron We construct the form of wavefunctions for the system with two, three, four, and five electrons (identical particles, fermion) using the Slater determinant. We assume that each electron has spin states

   ,    .

We also assume the orbital state: a  n  1 b  n  2 , c  n  3 .., where n is the principal quantum number. It is reasonable to consider that the ground state can be expressed by electrons occupied by

a  , a  taking into account of the Pauli’s exclusion principle. For the system with three electrons, the ground state can be expressed by

a  , a  , b 

For the system with the four electrons, the ground state can be expressed by

a  , a  , b  , b 

For the system with the five electrons, the ground state can be expressed by

a  , a  , b  , b  , ` c 

(a) Two particles state For the two states,

a  a  , and a  a  we have the Slater determinant

1 a11 a11 1    a1a2 (12  12 ) 2 a22 a22 2

Identical particles 15 9/3/2017 where index 1, 2 denotes the particle 1 and particle 2. Clearly, the state vector is antisymmetric under the exchange of two particles.

(b) Three particles state For the three states,

a  a  , a  a  , and b  b  we have the Slater determinant

a11 a11 b11 1   a  a  b  6 2 2 2 2 2 2 a33 a33 b33 1  [a a (     )b  a a (    )b   a a (     )b  ] 6 2 3 2 3 2 3 1 1 1 3 1 3 1 3 2 2 1 2 1 2 1 2 3 3

This state vector is antisymmetric under the exchange between any two particles (such that 1  2 ).

(c) Four particle state For the four states,

a  a  , a  a  , b  b  , and b  b  we have the Slater determinant

a11 a11 b11 b11 1 a  a  b  b    2 2 2 2 2 2 2 2 24 a33 a33 b33 b33

a4 4 a44 b44 b4 4

 b1b2 (12  1 2 )a3a4 (34  34 )

 b2b3 (23  23 )a1a4 (14  1 4 )

 b2b4 (24  2 4 )a1a4 (14  14 )

 b1b3 (13  13 )a2a4 (24  24 )

 b1b4 (14  14 )a2a3 ( 23  23 )

 b3b4 (34  34 )a1a2 (12  1 2 )

This state vector is antisymmetric under the exchange between any two particles (such that 1  2 ), but symmetric under the two types exchange (such that 1  2 and 3  4 )

(d) Five particle state For the five states Identical particles 16 9/3/2017

a  a  , a  a  , b  b  , b  b  , c  c  we have the Slater determinant

a11 a11 b11 b11 c11

a22 a22 b22 b22 c22 1   a  a  b  b  c  120 3 3 3 3 3 3 3 3 3 3 a44 a44 b44 b44 c44

a55 a55 b55 b55 c55 1  [a a (     )b b (     )c   ...] 120 1 2 1 2 1 2 3 4 3 4 4 3 5 5

((Mathematica))

Identical particles 17 9/3/2017 a1 1 a1 1 Clear "Global` " ;A1 ; a2 2a2 2 Det A1 FullSimplify a1 a2 2 1 1 2

a1 1a11b11 B1 a2 2a22b22 ; a3 3a33b33 Det B1 FullSimplify a1 a3 b2 a2 b3 2 3 1 a2 a3 b1 a1 b3 1 3 2 a3 a2 b1 a1 b2 1 2 3

a1 1a11b11b11 a2 2a22b22b22 C1 ; a3 3a33b33b33 a4 4a44b44b44 Det C1 FullSimplify b2 a3 a4 b1 2 1 1 2 4 3 3 4 a1 a4 b3 3 2 2 3 4 1 1 4 a3 b4 3 1 1 3 4 2 2 4 a2 a4 b1 b3 3 1 1 3 4 2 2 4 b4 a3 b1 3 2 2 3 4 1 1 4

a1 b3 2 1 1 2 4 3 3 4

______REFERENCES S. Tomonaga Angular momentum and spin (Misuzo Syobo, Tokyo, 1989) [in Japanese]. J.J. Sakurai Modern Quantum Mechanics, Revised Edition (Addison-Wesley, Reading Massachusetts, 1994). L.D. Landau and E.M. Lifshitz, Quantum Mechanics (Pergamon Press, Oxford, 1977). John S. Townsend , A Modern Approach to Quantum Mechanics, second edition (University Science Books, 2012). David H. McIntyre, Quantum Mechanics A Paradigms Approach (Pearson Education, Inc., 2012). C. Cohen-Tannoudji, B. Du, and F. Laloe, Quantum Mechanics vol. 2 (John Wiley & Sons, 1977).

APPENDIX-I ((Townsend textbook))

Identical particles 18 9/3/2017 For a symmetric spatial state under the exchange of two particles, we have

1   dr' dr"Sˆ r',r" r',r" Sˆ   dr' dr" r',r" r',r"  S 2   S   S and for an anti-symmetric spatial state under the exchange of two particles, we have

1   dr' dr"Aˆ r',r" r',r" Aˆ   dr' dr" r',r" r',r"  A 2   A   A where

r',r"  r' r" 1 2

1 1 Sˆ  (1ˆ  Pˆ ) , Aˆ  (1ˆ  Pˆ ) 2 12 2 12

((Proof))

1   dr' dr"Sˆ r',r" r',r" Sˆ  S 2   S 1 1  dr' dr" (1ˆ  Pˆ ) r',r" r',r" (1ˆ  Pˆ ) 2   2 12 12 S 1  dr' dr"(1ˆ  Pˆ ) r',r" r',r" (1ˆ  Pˆ ) 4   12 12 S 1  dr' dr"( r',r"  r",r' r',r"  2   S 1  dr' dr"( r',r" r',r"   r",r' r',r"  ) 2   S S 1  dr' dr"( r',r" r',r"   r",r' r",r'  ) 2   S S  dr' dr"( r',r" r',r"    S where we use the relation

ˆ r',r"  S  r',r" P12  S  r",r'  S .

Similarly,

Identical particles 19 9/3/2017 1   dr' dr"Aˆ r',r" r',r" Aˆ  A 2   A 1  dr' dr"(1ˆ  Pˆ ) r',r" r',r" (1ˆ  Pˆ ) 4   12 12 A 1  dr' dr"( r',r"  r",r' r',r"  A 2   1  dr' dr"( r',r" r',r"   r",r' r',r"  ) 2   A A 1  dr' dr"( r',r" r',r"   r",r' r",r'  ) 2   A A  dr' dr"( r',r" r',r"    A where we use the relation

ˆ r',r"  S  r",r' P12  S  r",r'  S

ˆ r',r"  A   r',r" P12  A   r",r'  A

ˆ ˆ P12  S   S , P12  A    A

______APPENDIX-II Property of permutation operator

ˆ 2 ˆ 1. P12  1

((Proof))

Pˆ 2  '  Pˆ '    ' 12 1 2 12 1 2 1 2

ˆ 2 ˆ ˆ leading to the relation P12  1. The operator P12 is its own inverse.

ˆ ˆ  ˆ 2. P12 is Hermitian operator: P12  P12

((Proof))

 ' Pˆ   '  (  ' )(  '  ) 1 2 12 1 2 1 2 1 2   , '  ',

From the definition

Identical particles 20 9/3/2017  ' Pˆ    '    ' Pˆ  ' * 1 2 12 1 2 1 2 12 1 2  [(   ' )( '  )]* 1 2 1 2 *  (  , '   ', )

   , '   ',

ˆ  ˆ This leads to the relation, P12  P12 .

ˆ ˆ  ˆ ˆ ˆ  ˆ 3. P12 is unitary operator: P12 P12  P12P12  1

((Proof))

ˆ  ˆ ˆ ˆ ˆ P12 P12  P12P12  1

ˆ ˆ  ˆ ˆ ˆ P12P12  P12P12  1

______APPENDIX-III Permutation operator (a)

1 2 3 4 P'   4 3 2 1 1 2 3 42 1 3 4     2 1 3 44 3 2 1 1 2 3 41 2 3 4     2 1 3 43 4 2 1

 P12P

1 2 3 43 4 2 1 P'    3 4 2 14 3 2 1 1 2 3 41 2 3 4     3 4 2 11 2 4 3

 PP34 leading to the expression in quantum mechanics;

ˆ ˆ ˆ ˆ PP12  P34P

(b)

Identical particles 21 9/3/2017 1 2 3 4 P'   3 1 4 2 1 2 3 43 2 1 4     3 2 1 43 1 4 2 1 2 3 41 2 3 4     3 2 1 44 1 3 2

 P13P

1 2 3 4 P'   3 1 4 2 1 2 3 44 1 3 2     4 1 3 23 1 4 2 1 2 3 41 2 3 4     4 1 3 21 2 4 3

 PP34 leading to the expression in quantum mechanics;

ˆ ˆ ˆ ˆ PP13  P34P

Identical particles 22 9/3/2017